Quantum Computation, Spectroscopy of Trapped Ions, and ...

4 downloads 0 Views 94KB Size Report
... Bollinger, R. Blatt,. D. Sullivan, and M. Young for helpful comments on the manuscript. ... J. Steinbach and C.C. Gerry, quant-ph/9806091. 16. See, for example ...
arXiv:quant-ph/9809028v1 10 Sep 1998

1

Quantum Computation, Spectroscopy of Trapped Ions, and Schr¨odinger’s Cat ∗ D.J. Wineland, C. Monroe, W.M. Itano, D. Kielpinski, B.E. King, C.J. Myatt, Q.A. Turchette, and C.S. Wood National Institute of Standards and Technology (NIST), Boulder, CO, 80303 We summarize efforts at NIST to implement quantum computation using trapped ions, based on a scheme proposed by J.I. Cirac and P. Zoller (Innsbruck University). The use of quantum logic to create entangled states, which can maximize the quantum-limited signal-to-noise ratio in spectroscopy, is discussed. 1. INTRODUCTION The invention by Peter Shor [1] of a quantum algorithm for factorizing large numbers has stimulated a host of theoretical and experimental investigations in the field of quantum information [2]. In the area of quantum computation, various schemes have been proposed to realize experimentally a model quantum computer [2]. In the ion storage group at NIST, we are trying to realize such a device based on the proposal by Cirac and Zoller [3]. In the Cirac-Zoller scheme, qubits are formed from two internal energy states, labeled | ↓i and | ↑i, of trapped atomic ions. If the ions are laser cooled in the same trap, they form a crystalline array whose vibrations can be described in terms of normal modes. The ground and first excited states of a selected mode can also form a qubit. This qubit can serve as a data bus, since the normal modes are a shared property of the ions. An individual ion in the array can be coherently manipulated and coupled to the selected normal mode by using focused laser beams [3]. A universal logic operation, such as a controlled-not (CN) logic gate between ion qubit i and ion qubit j, is accomplished by (1) mapping the internal state of qubit i onto the selected motional qubit, (2) performing a CN between the motional qubit and qubit j, and (3) mapping the motional qubit state back onto qubit i. Each of these steps has been accomplished in the NIST experiments with a single ion [4,5]. We are currently devoting efforts to: (1) scaling quantum logic operations to two or more ions (Sec. 5), (2) applying quantum logic to study fundamental measurement problems on EPR and GHZ-like states, and (3) applying quantum logic to fundamentally improve the signal-to-noise ratio (SNR) in spectroscopy and atomic clocks. In this paper we briefly discuss this last application. We are aware of similar efforts to implement trapped-ion quantum logic at IBM, Almaden; Innsbruck University; Los Alamos National Laboratory; Max Planck Institute, Garching; and Oxford University. ∗

Contribution of NIST; not subject to U.S. copyright

2 2. ENTANGLED STATES FOR SPECTROSCOPY A collection of atoms (neutral or charged) whose internal states are entangled in a specific way can improve the quantum-limited SNR in spectroscopy. This application of quantum logic to form entanglement is useful with a relatively small number of atoms and logic operations. For example, for high-accuracy, ion-based frequency standards [6], a relatively small number of trapped ions (L ≤ 100) appears optimum due to various experimental constraints; with L = 10 − 100, a significant improvement in performance in atomic clocks could be expected. In contrast, factoring a number which cannot easily be factored on a classical computer would require considerably more ions and operations. In spectroscopy experiments on L atoms, in which the observable is atomic population, we can view the problem in the following way using the spin-1/2 P analog for two-level atoms. The total angular momentum of the system is given by J = Li=1 Si , where Si is the spin of the ith atom (Si = 1/2). The task is to measure ω0 , the frequency of transitions between the | ↓i and | ↑i states, relative to the frequency ωR of a reference oscillator. We first prepare an initial state for the spins. Typically, spectroscopy is performed by applying (classical) fields of frequency ωR for a time TR according to the method of separated fields by Ramsey [7]. We assume the same field amplitude is applied to all atoms (the phases might be different) and that the maximum value of TR is fixed by experimental constraints (Sec. 3). After applying these fields, we measure the final state populations; for example, the number of atoms L↓ in the | ↓i state. In trapped-ion experiments, this has been accomplished through laser fluorescence detection with nearly 100% efficiency, which we assume here (see the discussion and references in Ref. [5]). In the spin-1/2 analog, measuring L↓ is equivalent to measuring the operator Jz , since L↓ = JI−Jz where I is the identity operator. The SNR (for repeated measurements) is fundamentally limited by the quantum fluctuations in the number of atoms which are observed to be in the | ↓i state. These fluctuations can be called quantum projection noise [8]. Spectroscopy is QL typically performed on L initially unentangled atoms (for example, Ψ(t = 0) = i=1 | ↓ii ) which remain unentangled after the application of the Ramsey fields. For this case, the imprecision in a determination of the frequency √ of the transition is limited by projection noise to the “shot noise” limit (∆ω)meas = 1/ LTR τ where τ ≫ TR is the total averaging time [8]. If the atoms can √ be prepared initially in particular entangled states, it is possible to achieve (∆ω)meas < 1/ LTR τ . In optics, squeezed states have been shown to improve the SNR in interferometers beyond the shot noise limit [9,10]. In 1986, Yurke [11] showed how particular entangled states, if they could be created, could be used as inputs to Mach-Zehnder interferometers to approach the Heisenberg limit of SNR. In 1991, Kitegawa and Ueda [12] showed how the Coulomb interaction between electrons in the two arms of an electron interferometer might be used to improve the SNR beyond the shot-noise limit. Because of the formal identity of Mach-Zehnder interferometers and Ramsey spectroscopy [13], similar ideas might be applied to the spectroscopy problem. Reference [13] showed how a Jaynes-Cummingstype coupling between trapped-ion internal states and a normal mode could be used to improve the SNR in spectroscopy beyond the shot-noise limit. The scheme in Ref. [13] has the advantage that the appropriate states can be generated by acting on all the ions at once (thus not requiring focused laser beams), but has the disadvantage that these states

3 are entangled with the motion, thereby requiring small motional decoherence. Reference [14] investigated the use of the generalized GHZ state, sometimes called the maximally entangled state, in spectroscopy. This state has the form ψmax

  1 iφ(t) | ↑i1 | ↑i2 · · · | ↑iL , = √ | ↓i1 | ↓i2 · · · | ↓iL + e 2

(1)

where φ(t) = φQ 0 − Lω0 t. After application of the Ramsey radiation, we measure the ˜ ≡ L Szi . The resulting signal gives the exact Heisenberg limit of SNR operator O i=1 √ ((∆ω)meas = 1/L TR τ where τ ≫ TR ) in spectroscopy (and interferometry). The state ψmax can be generated in a straightforward way by the application of L CN gates [3]. An alternative method was suggested in Ref. [14] and in Refs. [5] and [15] methods to generate ψmax with a fixed number of steps (independent of L) are discussed. For all of these methods, the the motion is entangled with internal states during the creation of ψmax , but is not entangled afterwards. Therefore, once ψmax is created, the motion can lose coherence without affecting the entanglement of the internal states. 2.1. Schr¨ odinger’s Cat As L becomes large and more macroscopic, states like ψmax become more like Schr¨odinger’s cat in that they represent coherent superpositions between widely separate regions of a large Hilbert space; for example, | ↑i1 | ↑i2 · · · | ↑iL ⇐⇒ “live cat;” | ↓i1| ↓i2 · · · | ↓iL ⇐⇒ “dead cat”. As has been emphasized in many discussions, as L becomes large the coherence between the two components of the cat becomes harder and harder to preserve [16]. This is apparent in Eq. (1) because if, for example, ω0 fluctuates randomly, the two components of ψmax will decohere relative to each other L times faster than for one ion (ψmax for L = 1). Trapped ions are interesting because it may be possible to make L very large without significant decoherence. This is the same property that makes trapped ions interesting as possible frequency standards. For example, in Refs. [17] and [18], coherence times for individual ions (L = 1) exceeding 10 minutes were obtained. 3. Applicability In the above, we have assumed that TR is fixed, limited by some independent experimental factor. This assumption is warranted in many trapped-ion atomic clock experiments, where, for example, we want to limit the heating that takes place with laser cooling radiation absent. (During application of the Ramsey fields the cooling radiation must be removed to avoid perturbing the clock states.) Additionally, we may want to lock a local oscillator to the atomic reference in a practical time [6,19], thereby limiting TR . However, the use of entangled states may not be advantageous, given other conditions. For example, Huelga, et al. [20] assume that the ions are subject to a certain dephasing decoherence rate (decoherence time less than the total observation time). In this case, there is no advantage of using maximally entangled states over unentangled states. The reason is that since the maximally entangled state decoheres L times faster than the states of individual atoms, when we use the maximally entangled state, TR must be reduced by a factor of L for optimum performance. Therefore, the gain from using the maximally entangled state is offset by the required reduced value of TR .

4 Reference [5] discusses another case of practical interest. In atomic clocks, the frequency of an imperfect “local” oscillator, whose radiation drives the atomic transition, is controlled by the atom’s absorption resonance. Depending on the spectrum of this oscillator’s frequency fluctuations (when not controlled) the use of entangled states may or may not be beneficial. 4. Implementations If we are able to create, with good fidelity, the state ψmax (Eq. (1)), how do we perform spectroscopy? First, we note that ψmax is the state we want after the first Ramsey π/2 pulse. Therefore, if we were to follow as closely as possible the Ramsey technique, we would take ψmax and apply a π/2 pulse of radiation at frequency ω0 to make the input state for the Ramsey radiation. However the first Ramsey π/2 pulse would only reverse this step; therefore, it is advantageous to take the creation of ψmax as the first Ramsey π/2 pulse. The second Ramsey pulse (after time TR ) can be applied directly with radiation at frequency ωR . The phase of this pulse (on each ion) must be fixed relative to the phases of the radiation used to create ψmax . In general, the relation between these phases and φ0 (Eq. (1)) will depend on the relative phases of the fields at the positions of each of the ˜ ∝ cos(L∆ωTR + φf ) where ∆ω ≡ ωR − ω0 ions [5,21]. This will lead to a signal S = hOi and where φf depends on all of these phases. ˜ as a function of TR , with ∆ω fixed. We can extract ω0 (relative to ωR ) by measuring hOi This can be further simplified by measuring the signal for two values of TR , TR2 ≫ TR1 , ˜ ≃ 0. Unfortunately, if the measured signal has a systematic bias as a function where hOi of TR , an error in the determination of ∆ω will result. This might happen, for example, if the ions heat up during application of the Ramsey radiation and a loss of signal occurs due to a reduced overlap between the ions and the laser used for fluorescence detection ˜ for two values of ωR , of the states. This problem could be overcome by measuring hOi ωR1 and ωR2 such that ωR1 − ω0 ≃ −(ωR2 − ω0 ) (determined by the above method), and two values of TR , TR1 ≪ TR2 . We then iterate the following steps: (1) we make ˜ R1 − ω0 )TR1 )i ≃ hO((ω ˜ R2 − ω0 )TR1 )i by adjusting the phase of the final π/2 pulse hO((ω to make φf → 0. This will take a negligible amount of time since TR1 ≪ TR2 . (2) We ˜ R1 − ω0 )TR2 )i ≃ hO((ω ˜ R2 − ω0 )TR2 )i by adjusting ωR1 and/or ωR2 to force make hO((ω ˜ has a systematic bias ωR1 − ω0 → −(ωR2 − ω0 ). This gives ω0 relative to ωR even if hOi as a function of TR . An alternative solution is suggested by Huelga, et al. [20]. After TR , instead of applying a π/2 pulse of radiation at frequency ωR , we apply the time-reversed sequence of operations which created ψmax . This has the advantage of cancelling out all of the CN phases that contribute to φ0 and maps the signal (∝ cos(L∆ωTR )) onto a single ion (whereupon Sz is measured for that ion). This also reduces the problem of detection efficiency to one ion rather than L ions. The disadvantage of this technique is that for large values of TR , the motional mode used for logic will, most likely, have to be recooled. This would require sympathetic cooling with the use of an ancillary ion which, to avoid the decohering effects of stray light scattering on the logic ions, might have to be another ion species [5]. A more serious limitation to the accurate determination of ω0 is that, in practice, ψmax will be realized only approximately and the state produced by the logic operations

5 will also be composed of states other than the | ↑i1 | ↑i2 · · · | ↑iL and | ↓i1 | ↓i2 · · · | ↓iL states; these other states will have a definite phase relation to the | ↑i1 | ↑i2 · · · | ↑iL and | ↓i1 | ↓i2 · · · | ↓iL states. Consequently, in general, the signal produced with either implemenation will be of the form

S=

L X

Cp cos(p∆ωTR + ξp ).

(2)

p=1

To accurately determine ∆ω, it will be necessary to Fourier decompose S. Since this will take more measurements, the advantages of using entangled states will be reduced. In spite of this, in some applications, it will be useful to determine changes in ω0 with respect to some external influence. For example, we might want to detect changes in ω0 caused by changes in an externally applied field. In this case, as long as |Cp | ≪ 1, for all p < L, we derive the benefits of entangled states (assuming the decoherence time is longer than TR /L) by measuring changes in S for a particular value of TR . 5. Experiments As usual, our enthusiasm for implementing these schemes far exceeds what is accomplished in the laboratory; nevertheless, some encouraging signs are apparent from recent experiments. In Ref. [22], all motional modes for two trapped ions have been cooled to the ground state. The non-center-of-mass modes are observed to be much less susceptible to heating, suggesting the use of these modes in quantum computation or quantum state engineering. In Ref. [21], we describe logic operations which enabled ψmax for L = 2 to be generated with modest fidelity (≃ 0.7). For small L, it is only necessary to differentially address individual ions to create ψmax and for L = 2, general logic can be realized even if the laser beams cannot be focused exclusively on the individual ions [21]. For general logic on more than two ions, two avenues are being pursued. For modest numbers of ions in a trap, the Cirac-Zoller scheme of individual addressing with the use of focused laser beams is the most attractive. Current efforts are devoted to obtaining sufficiently strong focusing to achieve individual ion addressing in a relatively strong trap where normal mode frequencies are relatively high (≃ 10 MHz) in order to maximize operation speed. Alternatively, general logic on many ions could be accomplished by incorporating accumulators [5], and using differential addressing on two ions at a time. This idea might be realized by scaling up a version of a linear ion trap made with lithographically deposited electrodes as we have recently demonstrated [16,23]. Concurrently, efforts are being devoted to the investigation (and hopefully, elimination) of mode heating [5] for different electrode surfaces and dimensions. 6. Acknowledgments We gratefully acknowledge the support of the U.S. National Security Agency, U.S. Army Research Office, and the U.S. Office of Naval Research. We thank J. Bollinger, R. Blatt, D. Sullivan, and M. Young for helpful comments on the manuscript.

6 REFERENCES 1. P.W. Shor, Proc. 35th Ann. Symp. Foundations of Computer Science, S. Goldwasser ed., IEEE Computer Society Press, New York, 1994, p. 124. 2. See, for example Proc. Royal Soc., Math., Phys. and Eng. Sci., 454 (1969) (1998); Fortschritte der Physik 46 (4-5) (1998). 3. J.I. Cirac and P. Zoller, Phys. Rev. Lett. 74, 4091 (1995). 4. C. Monroe, D.M. Meekhof, B.E. King, W.M. Itano, and D.J. Wineland, Phys. Rev. Lett. 75, 4714 (1995). 5. D.J. Wineland, C.R. Monroe, W.M. Itano, D. Leibfried, B.E. King, and D.M. Meekhof, NIST J. Research 103 (3), 259 (1998). (available at http://nvl.nist.gov/pub/nistpubs/jres/jres.htm) 6. D.J. Berkeland, J.D. Miller, J.C. Bergquist, W.M. Itano, and D.J. Wineland, Phys. Rev. Lett. 80, 2089 (1998). 7. N.F. Ramsey, Molecular Beams (Oxford Univ. Press, London, 1963). 8. W. M. Itano, J. C. Bergquist, J. J. Bollinger, J. M. Gilligan, D. J. Heinzen, F. L. Moore, M. G. Raizen, and D. J. Wineland, Phys. Rev. A47, 3554 (1993). 9. C.M. Caves, Phys. Rev. D23, 1693 (1981). 10. M. Xiao, L. -A. Wu, and H. J. Kimble, Phys. Rev. Lett. 59, 278 (1987). 11. B. Yurke, Phys. Rev. Lett. 56, 1515 (1986); B. Yurke, S.L. McCall, and J.R. Klauder, Phys. Rev. A33, 4033 (1986). 12. M. Kitagawa and M. Ueda, Phys. Rev. Lett. 67, 1852 (1991); M. Kitagawa and M. Ueda, Phys. Rev. A47, 5138 (1993). 13. D.J. Wineland, J. J. Bollinger, W. M. Itano, F. L. Moore, and D. J. Heinzen, Phys. Rev. A46, R6797 (1992); D.J. Wineland, J. J. Bollinger, W. M. Itano, and D. J. Heinzen, Phys. Rev. A50, 67 (1994). 14. J.J. Bollinger, D. J. Wineland, W. M. Itano, and D. J. Heinzen, Phys. Rev. A54, R4649 (1996). 15. J. Steinbach and C.C. Gerry, quant-ph/9806091. 16. See, for example, S. Haroche, Physics Today 51 (7), 36 (1998). 17. J.J. Bollinger, D. J. Heinzen, W. M. Itano, S. L. Gilbert, and D. J. Wineland, IEEE Trans. on Instrum. and Measurement 40, 126 (1991). 18. P.T.H. Fisk, M.J. Sellars, M.A. Lawn, C. Coles, A.G. Mann, and D.G. Blair, IEEE Trans. Instrum. Meas. 44, 113 (1995). 19. J.J. Bollinger, J.D. Prestage, W.M. Itano, and D.J. Wineland, Phys. Rev. Lett. 54, 1000 (1985). 20. S.F. Huelga, C. Macchiavello, T. Pellizzari, A.K. Ekert, M.B. Plenio, and J.I. Cirac, Phys. Rev. Lett. 79, 3865 (1997). 21. Q.A. Turchette, C.S. Wood, B.E. King, C.J. Myatt, D. Leibfried, W.M. Itano, C. Monroe, and D.J. Wineland, submitted (quant-ph/9806012). 22. B.E. King, C.S. Wood, C.J. Myatt, Q.A. Turchette, D. Leibfried, W.M. Itano, C. Monroe, and D.J. Wineland, Phys. Rev. Lett. 81, 1525 (1998). 23. C. Myatt, et al., NIST Ion Storage Group, unpublished.